Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Results in Physics 47 (2023) 106334

Contents lists available at ScienceDirect

Results in Physics
journal homepage: www.elsevier.com/locate/rinp

Dielectric relaxations and low dissipation factor with excellent temperature


stability of Ti1-x(Co1/3Ta2/3)xO2 ceramics
Theeranuch Nachaithong a, b, Narong Chanlek c, Pairot Moontragoon a, b,
Pornjuk Srepusharawoot a, b, Prasit Thongbai a, b, *
a
Giant Dielectric and Computational Design Research Group (GD–CDR), Department of Physics, Faculty of Science, Khon Kaen University, Khon Kaen 40002, Thailand
b
Institute of Nanomaterials Research and Innovation for Energy (IN-RIE), Khon Kaen University, Khon Kaen 40002, Thailand
c
Synchrotron Light Research Institute (Public Organization), 111 University Avenue, Muang District, Nakhon Ratchasima 30000, Thailand

A R T I C L E I N F O A B S T R A C T

Keywords: Low dissipation factor (tanδ) and excellent temperature stability of high dielectric permittivity are obtained in
Giant/Colossal permittivity Ti1-x(Co1/3Ta2/3)xO2 ceramics, which were prepared using a mixed oxide method. The unit cell of the rutile TiO2
Grain boundary barrier layer capacitor structure expanded due to relatively larger ionic radii of the co–dopants. The dopants homogeneously appeared
Impedance spectroscopy
in the microstructure of the TiO2 ceramics. Highly dense sintered–ceramics without second phase exhibit low
TiO2
tanδ of 0.015–0.043 with very high dielectric permittivities of 7.0–9.1 × 104 at 1 kHz. Notably, the temperature
Temperature coefficient
coefficient in the range of ±15 % was obtained in the temperature range of − 60–200 ◦ C. The electrically het­
erogeneous microstructure was studied using impedance spectroscopy. The contributions of extrinsic and
intrinsic factors were separated by the dielectric relaxations in low and high–frequency ranges, respectively. The
dielectric relaxations were well fitted using a modified Cole-Cole model. A low–frequency relaxation was well
described based on the Maxwell-Wagner relaxation model. A high permittivity was originated from the intrinsic
(i.e., defect dipoles) and extrinsic (internal barrier layer capacitor, IBLC) effects. A low tanδ was primarily caused
by the IBLC effect associated with a high resistance of the insulating part and defect dipoles.

Introduction Before the remarkable dielectric properties of these materials were


discovered, CaCu3Ti4O12 (CCTO), which is in the ACu3Ti4O12 (A is
TiO2 is a very abundant material in nature and has many applications divalent or average divalent ions), is the most famous giant dielectric
for chemical, physical, optical, and electrical applications. For a ceramics. This material family showed ε′ ~ 103–105 values; however,
dielectric application, TiO2 was used classified as medium permittivity their tanδ values were higher than the standard applications [10–15].
materials with dielectric permittivity of ε′ ~ 90–200, which was used in Furthermore, it was likely that Δε′ < ±15 % was obtained only in a
a high–power capacitor [1]. In recent years, the material development narrow temperature range (− 55–100 ◦ C) [13,16–20]. Thus, it is hardly
intention of rutile TiO2 has transferred from a medium permittivity to accomplish the significantly improved DPs from the CCTO–based
material to its use as a high permittivity material with high ε′ ~ 104–105. materials. Although a low tanδ value can be obtained in BaTiO3–based
A large ε′ value of TiO2 is referred to as a giant dielectric response. The materials, the Δε′ value is usually high near a curie temperature [21].
dielectric properties (DPs) of TiO2 can be significantly changed by Furthermore, these materials are complex oxides, which are difficultly
doping process with alivovalent and/or isovalent ions [2,3]. Usually, prepared in a large scale compared to those of simple oxides (e.g., CuO,
enhanced DPs with ε′ ~ 104 and low tanδ < 0.1 are obtained, while the NiO, and TiO2). Nevertheless, a tanδ value of NiO–based oxide was very
temperature coefficient for X–R standard applications (Δε′ < ±15 %) is large even through its giant DPs can be achieved by co–doping with
achieved in the range of − 55–180 ◦ C. Therefore, the DPs of TiO2 have aliovalent ions [22,23]. On the other hand, TiO2–based ceramics showed
increasingly become the focus of research activity [2,4–9]. Accordingly, a largely increased ε′ on the order of 104 and greatly reduced tanδ ≈
TiO2 has a potential application in multilayer ceramic capacitors 10− 2, which were performed by using aliovalent dopants of In3+ and
(MLCCs) and related electronic devices [5]. Nb5+ (InNbTO) [2]. TiO2 is a simple oxide and abundant material in

* Corresponding author at: Giant Dielectric and Computational Design Research Group (GD–CDR), Department of Physics, Faculty of Science, Khon Kaen Uni­
versity, Khon Kaen 40002, Thailand.
E-mail address: pthongbai@kku.ac.th (P. Thongbai).

https://doi.org/10.1016/j.rinp.2023.106334
Received 31 January 2023; Received in revised form 23 February 2023; Accepted 2 March 2023
Available online 6 March 2023
2211-3797/© 2023 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-
nc-nd/4.0/).
T. Nachaithong et al. Results in Physics 47 (2023) 106334

nature. crystal structure and phase composition in the sintered specimens. As


It was reported that the DPs the InNbTO materials were changed illustrated in Fig. 1(a), the XRD patterns provide the phase composition
with the concentration of the dopant. The dielectric response increases data with a single phase of rutile TiO2 (JCPDS 21-1276) only for all the
as the dopants changed from 0.05 to 10 %, while a low-frequency tanδ samples. A slight difference in the XRD patterns of the CoTTO and TiO2
slightly reduced. Furthermore, the dielectric response was also weakly was found. To further clarify the possibility of the substitution of the
dependent on temperature and frequency. The comprehensive details of dopants into the TiO2 lattice, the lattice parameters were calculated. As
the investigations of the DPs for many codoped TiO2 materials, such as shown in Fig. 1(b), obviously, both the a and c values of the unit cell of
the Zn2+/Nb5+, La3+/Ta5+, Ga3+/Nb5+, Tm3+/Nb5+, Sc3+/Nb5+, Al3+/ the CoTTO samples are larger than that of the undoped TiO2. Both values
Nb5+, Al3+/Ta5+, Zr4+/Sb5+, and etc. [3,24–31], were reported. Most of also slightly increased with increasing x from 0.5 to 5 %. Both the ionic
them showed significantly improved DPs, as mentioned above. Form this radii of Ta5+ (r6 = 0.64 Å) and Co2+ (r6 = 0.65 Å) are larger than the Ti4+
point of views, codoped TiO2 ceramics can be classified as a promising (r6 = 0.605 Å) [35]. Therefore, the a and c values may be enlarged by the
material group with high potential for use in MLCCs and related devices. substitution of the dopants, which was confirmed by linearly enlarged
For future application in a high technology, a low Δε′ (%) must be cell volume of TiO2 with increasing the co–doping concentration [inset
extended to a wide temperature range (>150 ◦ C) as far as possible. of Fig. 1(b)]. Considering a single that was found in all the samples and
Unfortunately, a low Δε′ (%) of the InNbTO ceramics and many codoped changes in cell parameters, both Co2+ and Ta5+ could be replaced Ti4+
TiO2 ceramics was obtained only in a temperature range of ≤150 ◦ C, in the rutile-TiO2 structure. It is important to note that a change in c
which may be one of most serious problems supporting a high tech­ value was larger than that of a value. He et al. reported that, according to
nology [32]. Hence, a new codoped TiO2 with improved Δε′ (%) over a the theoretical calculation and experimental results, the lowest energy
wide temperature range is important to seek. In addition, the important configuration of the InNbTO structure was obtained when the dopants
cause causes of the observed DPs must be clarified. Electric dipoles from located along the c axis, resulting in a decreased c/a ratio. Thus, in this
complex defect clusters and space changes at the internal interfaces may work, it is possible that the Co2+ and Ta5+ doping ions preferred to
have impacts on the overall DPs in different temperatures locate on the Ti4+ sites along the c axis.
[2,4,24–26,33,34]. Each proposed model is reasonable from different To further characterize using other techniques and study the
point of views, i.e., different temperature and frequency ranges. dielectric properties, the valent states of all elements in the 1.0 %CoTTO
Therefore, the objectives of this research are to seek a new codoped TiO2 ceramic were examined, as displayed in Fig. S1 (supplementary infor­
that can exhibit improved DPs in all parameters and to ascribe the mation). The minority of Ti3+ can be induced by doping with Ta5+
important causes of the dielectric responses. (Fig. S1(a)), as reported in previous papers [2,24,26,36]. Furthermore,
In this work, we synthesized a new codoped TiO2 system, i.e., Co2+ oxygen vacancies (VO⋅⋅ ) can be detected due to the charge compensation
and Ta5+, employing a mixed oxide (MO) method. The crystalline for the replacement of Ti4+ with Co2+ or Co3+ (Fig. S1(b)), while the
structure and microstructure were studied. Low tanδ values of presence of Ta5+ was confirmed. Details of the mechanisms of the
0.015–0.043 with very high ε′ of 7.0–9.1 × 104 were observed. Notably, induced Ti3+ and VO⋅⋅ were explained in the supplementary information.
a low temperature coefficient was achieved over the temperature range
of − 60–210 ◦ C. The interfacial polarization and defect dipoles were
suggested to be the important causes of the DPs.

Experimental details

(Co1/3Ta2/3)xTi1-xO2 (CoTTO), where x = 0.005, 0.01, and 0.05,


powders were synthesized. These compositions are referred to as the 0.5
%CoTTO, 1.0 %CoTTO, and 5 %CoTTO, respectively. The starting raw
oxides, purchased from Sigma-Aldrich, consisted of rutile-TiO2 (>99.9
%), CoO (99.9 % purity), and Ta2O5 (99.99 %). Details of the experi­
mental steps for preparing codoped TiO2 ceramics are given in the our
previous works [26]. The calcination temperature was 900 ◦ C for 8 h.
The powders were further processed by pressing them into pellets, fol­
lowed by sintering at 1400 ◦ C for 5 h with heating rate of 2 ◦ C/min.
A field–emission scanning electron microscopy (FE–SEM) with
energy–dispersive X-ray analysis (EDS) (HITACHI SU8030, Japan) and
SEM–mapping technique, X-ray diffraction (XRD, PANalytical, EMPY­ a
REAN), A UV–vis Raman spectrometer (Horiba Jobin–Yvon T64000), c
and X-ray photoelectron spectroscopy (XPS) (PHI5000 VersaProbe II,
ULVAC–PHI, Japan) were employed to systematically characterize the
sintered specimens. Comprehensive details of each technique are pro­
a Å

c Å

vided in our previous published work [26].


Silver paint was selected as an electrode material. After painting Ag
V Å3

on both surfaces, the samples were heated at 600 ◦ C for 0.5 h. The
capacitance (Cp)–dissipation factor (or tanδ) mode was selected for the
dielectric measurement. The ε′ values were calculated using the
measured Cp values. The measurement was performed as a function of
frequency (40–107 Hz) and temperature (− 60–210 ◦ C) by employing a
KEYSIGHT E4990A Impedance Analyzer. Vrms = 500 mV was used
throughout the dielectric resting.
Fig. 1. (a) XRD patterns of the sintered 0.5 %CoTTO, 1.0 %CoTTO, and 5.0 %
Results and discussion CoTTO compared to that of the undoped TiO2. (b) Relationship of lattice pa­
rameters and doping levels; inset shows co-doping concentration dependence of
In this research work, we started with the characterization of the cell volume for CoTTO ceramics compared to that of the undoped TiO2.

2
T. Nachaithong et al. Results in Physics 47 (2023) 106334

It is worth noting that the XPS peak of Co cannot be detected in the XPS performed. Fig. 3(a) displays the Raman peaks of a pure TiO2 and all
spectrum of the 1.0 %CoTTO ceramic. For CaCu3-xCoxTi4O12 ceramics, CoTTO specimens. Four peaks of the Raman spectra confirm the rutile
the co–existence of Co2+/Co3+ was detected when the Co doping con­ TiO2 structure, which are similar to many co–doped TiO2 materials
centration was increased from x = 1.0 to 2.0. However, only Co2+ can [3,36,42]. Only the strongest A1g and Eg peaks were focused because the
be detected when x = 0.05 [37]. In this work, it is possible that a small A1g and Eg modes were used to describe the variations in O–Ti–O bonds
portion of Co3+ might also be formed when the co–doping concentration and VO⋅⋅ , respectively [42]. Thus, both modes might relate to the DPs of
increased. Nevertheless, both Co2+ and Co3+ can cause the formation of co–doped TiO2 materials [2–4]. The correlation between the Raman
VO⋅⋅ in the rutile–TiO2 structure, which was similar to other TiO2 codoped shifts of these two modes and doping level is illustrated in Fig. 3(b).
with acceptor (+2/+3)/donor (+5) dopants [2–4,26,38]. On the other Obviously, the Raman peak of Eg mode appeared at lower wave numbers
hand, the co–existence of Co2+/Co3+ in CCTO ceramics caused an in­ as the codoped concentration increased from x = 0 to 5 %. Nevertheless,
crease in the giant DPs due to the enhanced mixed–valent structure of
Co2+/Co3+. This might also be responsible for the CoTTO ceramics.
Usually, the DPs of ceramics and their microstructures have a close
relation together [12,18,39]. Thus, the morphologies of the sintered
specimens were focused. As revealed in Fig. 2, no pores are observed in
the microstructure for all samples even at the grains and grain bound­
aries. The existence of all elements in the CoTTO ceramics were
confirmed using the EDS technique, as shown in Fig. S3(a) of supple­
mentary information. We also found that the dopants were well
dispersed throughout the microstructure using the SEM–mapping tech­
nique, as shown in Fig. S3(b). The average grain sizes (G) of all the
CoTTO samples slightly changed with G values of 10.6 ± 3.5, 7.3 ± 2.3,
and 10.9 ± 4.0 μm for the 0.5 %CoTTO, 1.0 %CoTTO, and 5.0 %CoTTO,
respectively. According to the previous works [26,40], the G of TiO2 was
greatly reduced by doping with Ta5+. Thus, the G should be significantly
reduced. However, a slightly changed G of the CoTTO ceramics was
observed, which was likely attributed to the dominant effect of Co2+
doping ions. The grain growth in polycrystalline ceramics is usually
determined by diffusion of atom or ions across the grain boundary.
Enhanced diffusion of ions can be primarily resulted from the formation
of liquid phase or increased point defect concentration (vacancies).
However, the melting point of CoO and eutectic temperature of
CoO–TiO2 are much higher than the sintering temperature of the CoTTO
ceramics [41]. Thus, the dominant effect of the Co2+ on the retained
grain size of the CoTTO ceramics may be attributed to the increase in
defect concentration. Doping TiO2 with Co2+ ions resulted in the crea­
tion of VO⋅⋅ , following:

(1)
TiO2
CoO ̅̅̅→ CoTi + VO⋅⋅ + OO

Thus, it is likely that during the sintering process, the enhanced


diffusion of VO⋅⋅ , which was confirmed to exist using the XPS technique
(Fig. S1(a)), is the key factor for contribution of the grain growth. The
restorative and driving forces acting on the grain boundary, which were
resulted from the Ta5+ and Co2+ dopants, were balanced, and hence the
G of the CoTTO ceramics slightly changed. In the other word, the
reduced G of the Ta5+–doped TiO2 was inhibited by the Co2+ doping
Fig. 3. (a) Raman spectra the sintered 0.5 %CoTTO, 1.0 %CoTTO, and 5.0 %
ions.
CoTTO compared to that of the undoped TiO2. (b) Relationship between the
To further confirm the existence of VO⋅⋅ the Raman spectroscopy was Raman shift (Eg and A1g models) and doping level.

Fig. 2. Surface morphologies of (a) 0.5 %CoTTO, (b) 1.0 %CoTTO, and (c) 5.0 %CoTTO using SEM technique.

3
T. Nachaithong et al. Results in Physics 47 (2023) 106334

according to Eq. (1), the primary cause of this observation is likely the Table 1
Co2+. The A1g peak position slightly shifted to a higher wave number, The ε′ and tanδ values at ~25 ◦ C and 1 kHz and temperature coefficient (Δε′ (%))
which was attributed to the variations in Ti–O stretching and O–Ti–O at the temperatures − 60 and 200 ◦ C.
bond bending due to the dopants. Parker et al. [43] demonstrated that x%CoTTO ε′ tanδ Δε′ (%)
the Eg peak of TiO2-x move backward from 447 to 437 cm− 1 as x = 0.02. − 60 ◦ C 200 ◦ C
The theoretical value of x in the 5.0 %CoTTO (in the cause of no self­
0.5 %CoTTO ~70,700 0.017 − 2.68 % 12.9 %
–charge compensation between Co2+–2Ta5+) is 0.017, the Eg peak
1.0 %CoTTO ~79,500 0.015 − 3.41 % 12.1 %
decreased from 447.5 (pure TiO2) to 435 cm− 1, which is consistent with 5.0 %CoTTO ~91,900 0.043 − 9.73 % 14.3 %
that reported by Parker et al. The Eg peak shifting to lower a wave
number showed the presence of more VO⋅⋅ in the CoTTO compared to that
of the undoped TiO2.
To explain the origin of enhanced DPs on the CoTTO ceramics, the
influence of the co–dopant concentration on the ε′ and tanδ at different
frequencies was studied. As illustrated in Fig. 4 and its inset, the ε′ in the
frequency range of 40–104 Hz of the CoTTO increases with co–dopant
content from 0.5 to 5.0 %. The step–like decrease in ε′ was observed at
the frequency higher than 104 Hz. This result was due to the relaxation
mechanism of a low–frequency dielectric response, which was likely due
to the extrinsic effect. tanδ in the frequency range of 40 to 106 Hz is
illustrated in the inset of Fig. 4. tanδ peak was observed in a high­
–frequency range, corresponding the appeared step–like decrease in ε′ .
The values of ε′ and tanδ at 1 kHz were listed in Table 1. Notably, all the
samples showed a very high ε′ of 7 × 104 with low tanδ values of
0.015–0.047. The excellent DPs of the CoTTO are comparable to those
reported in the (Ga3+/Ta5+) [26], (Ga3+/Nb5+) [34], (Zn2+/Nb5+) [44],
(Sc3+/Nb5+) [25], (In3+/Nb5+) [2,32,33], and (Al3+/Nb5+) [42] co- f
doped TiO2.
Fig. 5. Frequency dependence of ε′ in the temperature range from − 60 to 0 ◦ C
According to the extrinsic effects (i.e., interfacial polarization at in­ for 5 %CoTTO. Symbol and solid curve are the experimental data and fitted
ternal interface and resistive outer surface layer), a very large dielectric curves fitted using Eq. (2), respectively. Inset shows the Arrhenius plots of τlow− f
response of ε′ ~ 104 is generally obtained in a low–frequency range, for all CoTTO.
especially when the temperature decreased [11]. The dielectric response
due to the intrinsic effect can occur at a relatively low–temperature
range [2,4,34]. It is important to note that the contribution of i-GBs
′ ′
Δεlow− f Δεhigh− f
ε* = (2)

(IBLC) or electrodes is usually very high [45]. Thus, the electrode effect
( )α + ( )β + ε∞
1 + iωτlow− f 1 + iωτhigh− f
must be separated from the IBLC effect. Generally, the electrode effect
on the dielectric response is dominant at low frequencies and high where Δεlow−

and Δεhigh−

are the dielectric strengths in low (extrinsic
f f
temperatures. In this work, we tried to separate the dielectric responses
effect) and high–frequency (intrinsic effect) ranges, respectively, which
between the extrinsic (IBLC) and intrinsic effects by represent the DPs at
are estimated to be the contributions of the dielectric responses in these
low temperatures from − 60 to − 10 ◦ C, as illustrated in Fig. 5. Unfor­
two ranges. We set ε∞ ≈ 250 [2], which is the contribution from the

tunately, at the lowest measured temperature of − 60 ◦ C, a large


dielectric response due to the intrinsic response was entirely observed. atomic (Pa ) and electronic (Pe ) polarizations in a rutile TiO2. τlow− f and
Thus, we fitted the data using the modified Cole–Cole equation, using τhigh− f are the relaxation times in low and high–frequency ranges,
two dielectric relaxation parts: respectively. The data in a low–frequency range was well fitted by the
first term of Eq. (2). The data in a high–frequency range was fitted in
accordance with the second and third terms using τhigh− f ≈ 1.0–0.71 ns
and β = 1, as the temperature increased from − 60 to 0 ◦ C with a given
activation energy for a high relaxation of Ea,high− f ~ 30 meV for co–do­
ped TiO2 [2,34]. As illustrated in Fig. 5, the dielectric data were well
fitted by Eq. (2). According to the fitted curves in a low–frequency range,
τlow− f values of all the samples at the selected temperatures were ob­
tained with α ≈ 0.80–0.87. As displayed in the inset of Fig. 5, the tem­
perature dependence of τlow− f followed the Arrhenius law,
( )
Ea
τlow− f = τ0 exp (3)
kB T

where T is absolute temperature and kB is Boltzmann constant, τ0 is a


constant. According to the fitted results, the activation energy (Ea )
values required for dielectric relaxation in a low–frequency range
(Ea,low− f ) of the 0.5 %CoTTO, 1.0 %CoTTO, and 5.0 %CoTTO can be
calculated and found to be 77, 91, and 113 meV, respectively. The Ea
increased as the co–doping content increased. It is observed that the Ea
values of the CoTTO ceramics are similar to those reported in CCTO
(93–109 meV [46,47]), Bi2/3Cu3Ti4O12 (91–95 meV [10,46]), Y2/
3Cu3Ti4O12 (90 meV [46]), and La2/3Cu3Ti4O12 (80 meV [46]). The DPs
Fig. 4. ε′ as a function of frequency at ~25 ◦ C for 0.5 %CoTTO, 1.0 %CoTTO,
of these materials were well described using the Maxwell–Wagner
and 5.0 %CoTTO; inset shows tanδ as a function of frequency at 30 ◦ C.

4
T. Nachaithong et al. Results in Physics 47 (2023) 106334

relaxation (MWR) model based on the interfacial polarization (Pi ) at the Nevertheless, as illustrated in Fig. 5, it is likely that the dielectric
internal interface (or internal barrier layer capacitor, IBLC). Thus, it is response due to the intrinsic effect in the CoTTO can give rise to the ε′
likely that the dielectric response in a low–frequency range was caused value of 2.0 × 104, as well as observed in other codoped TiO2 systems
by the extrinsic effect due to Pi . [2,4,38]. Many research works believed such a dielectric response was
As shown in the inset of Fig. 5, τlow− f of the CoTTO ceramics increases attributed to the defect dipoles [2–4,38,44]. According to Eqs. (1) and
with increasing the codoping concentration, corresponding to their (S1–S3) in supplementary information, many defect dipoles can be
tanδ–peak positions [inset of Fig. 4] and the frequency at which the formed, i.e., Ti4+ − e− − VO⋅⋅ − Ti4+ − e− , VO⋅⋅ − VTi − VO⋅⋅ ,Ta⋅Ti − Ti4+ − e− , and
step–like decrease in ε′ appeared. When the average τlow− f was short­ CoTi − VO⋅⋅ . Furthermore, the complex defect clusters of
ened, the polarization can be occurred in a higher frequency range. CoTi − VO⋅⋅ +2Ta⋅Ti +2TiTi may also formed and have a great impact on the
According to the MWR model [10], it was theoretically demonstrated overall DPs, which can be active even at a high− frequency range and at
that the τlow− f was dependent on the semiconducting grain (semi-G) low temperatures. These defect dipoles were referred to as the electro­
resistance (Rg) and insulating grain boundary (i–GBs) capacitance (Cgb), n− pinned defect dipoles (EPDDs), which can be referred to as the
intrinsic effect. A very large dielectric response with ε′ ~ 104 can be
τlow− f ≈ Rg Cgb (4) extended into a high− frequency region by this way. Thus, the signifi­
To describe the dielectric relaxation behavior, an impedance spec­ cantly decrease in the ε′ associated with the frozen of polaron [49]. It is
troscopy was performed. Generally, the electrical responses of the sem­ worth noting that EPDDs were suggested to be the important cause of the
i− G and i− GBs are determined by small and large semicircular arcs, giant DPs in some codoped TiO2 systems [2,3,38]. The formation of
respectively [10,39]. If a small arc cannot be observed, the electrical EPDDs was dependent on the ionic size of the acceptor dopants [34],
response of semi− G can also be determined by a nonzero intercept on the which were formed in the In3+/Nb5+ and In3+/Ta5+ codoped TiO2
Z′ -axis [39]. Fig. 6 illustrates impedance complex plane (Z*) plots of all systems [2,4] due to a relatively large ionic radius of In3+ (r6 = 0.80 Å)
the CoTTO at 30 ◦ C. Only parts of a large arc were observed in all [35]. When the ionic radius of the acceptor dopant was not large enough
CoTTO, while small semicircular arcs were observed in the 1 %CoTTO or smaller than that of the host Ti4+ (e.g., Al3+ (r6 = 0.535 Å) or Ga3+ (r6
and 5 %CoTTO. As shown in the inset of Fig. 6, a nonzero intercept was = 0.62 Å)), the EPDDs were not created in the Al3+/Nb5+ [42] and
observed in the 0.5 %CoTTO. These results clearly indicated that semi-G Ga3+/Nb5+ [34] codoped TiO2 systems. Considering the ionic radius of
and i-GBs were produced in the TiO2 codoped with Co2+ and Ta5+. Co2+ (r6 = 0.65 Å), it is possible that the EPDDs can be formed in the
Accordingly, Pi can be induced under al applied electric field, giving rise CoTTO ceramics, which can contribute to the total dielectric response.
a large dielectric response in a low-frequency range. This is the first Besides very high ε′ and low tanδ values, the temperature depen­
reason proving the extrinsic effect. Considering a small arc and nonzero dence of the DPs is also one of the key factors for considering a dielectric
intercept in Z* plots, Rg of the CoTTO increased with increasing material for applications. The temperature coefficient of the ε′ at 1 kHz
codoping content. As shown in Fig. 4, the ε′ value below 104 Hz (Δε′ (%)) at any temperatures was calculated. As shown in Fig. 7,
increased as the codoping concentration increased, indicating the Notably, the Δε′ (%) values of all the CoTTO samples were less than ±15
enhanced Cgb value. From Eq. (4), the increased τlow− f was well described % in the temperature range from − 60 to 200 ◦ C. Usually, the Δε′ (%)
by the MWR model via the increased Rg and Cgb values. This is the values of many CGDO materials rapidly increased as the temperature
second reason for confirming the contribution of extrinsic effect in the increased higher than 150 ◦ C. As summarized in Table 1, at 200 ◦ C, the
CoTTO. The surface layer effect in the CoTTO can be ignored because Δε′ (%) values of the 0.5 %CoTTO, 1.0 %CoTTO, and 5.0 %CoTTO were
both sides of the outer surface layer of the CoTTO were removed before 12.0, 12.1, and 14.3 %, respectively. The Δε′ (%) values of the CoTTO
making electrodes for the dielectric measurement. Therefore, the IBLC can meet the standard values for the X7R, X8R, and X9R capacitors. The
was the primary effect on the low− frequency dielectric response. ε′ and tanδ values at ~25 ◦ C and 1 kHz of the 1.0 %CoTTO ceramic were
It is worth noting that a low-frequency relaxation due to the extrinsic compared to those of other giant dielectric oxides. As summarized in
effect in the CoTTO is similar to those observed in conventional giant Table 2, it is suggested that the CoTTO ceramics can be considered as
dielectric oxides (CGDO) [10,22,39,48]. However, the totally different one of the most interesting giant dielectric oxides. A low tanδ was pri­
behaviors between them are relatively high− frequency responses. The marily caused by a very large resistance of the i− GBs, as shown in Fig. 6.
dielectric response due to the intrinsic effect in CGDO usually give rise to Furthermore, confinement of free electrons in defect clusters can also be
the ε′ value of ~102 in a high− frequency range and low temperatures. a cause of a low tanδ.

semi

semi

Fig. 6. Z* plots at 30 ◦ C of all CoTTO; inset reveals an expanded view near the
origin, showing the nonzero intercept on Z′ -axis for the 0.5 %CoTTO. Fig. 7. Temperature coefficient of ε′ at 1 kHz for all CoTTO.

5
T. Nachaithong et al. Results in Physics 47 (2023) 106334

Table 2 org/10.1016/j.rinp.2023.106334.
The ε′ and tanδ values at ~25 ◦ C and 1 kHz of the 1.0 %CoTTO compared to
those of other giant dielectric oxides. References
Giant dielectric oxides ε′ tanδ References
[1] Moulson AJ, Herbert JM. Electroceramics: materials, properties, applications. 2nd
Co2+/Ta5+ codoped TiO2 ~7–9 × 0.015–0.043 In this ed. West Sussex; New York: Wiley; 2003.
104 work [2] Hu W, Liu Y, Withers RL, Frankcombe TJ, Norén L, Snashall A, et al. Electron-
Zn2+/Nb5+ codoped TiO2 ~3 × 104 0.05 [44] pinned defect-dipoles for high-performance colossal permittivity materials. Nat
Mg2+/Ta5+ codoped TiO2 >7 × 103 0.002 [38] Mater 2013;12(9):821–6.
In3+/Nb5+ codoped TiO2 ~6 × 104 0.02 [2] [3] Cheng X, Li Z, Wu J. Colossal permittivity in ceramics of TiO2 co-doped with
Ga3+/Nb5+ codoped TiO2 ~104–105 ~0.05–0.1 [34] niobium and trivalent cation. J Mater Chem A 2015;3(11):5805–10.
Li+/Ti4+ codoped NiO ~2 × 104 ~0.4 [22] [4] Dong W, Hu W, Frankcombe TJ, Chen D, Zhou C, Fu Z, et al. Colossal permittivity
Cu3Ta2Ti2O12 ~8 × 104 ~1.2 [50] with ultralow dielectric loss in In + Ta co-doped rutile TiO2. J Mater Chem A 2017;
5(11):5436–41.
CCTO ~104 ~0.1 [51]
[5] Wang Y, Jie W, Yang C, Wei X, Hao J. Colossal permittivity materials as superior
NaCu3Ti3SbO12 ~103–104 ~0.02–0.12 [52]
dielectrics for diverse applications. Adv Funct Mater 2019;29:1808118.
Na1/3Cd1/3Y1/3Cu3Ti4O12 ~4 × 104 0.065 [17]
[6] Zhou X, Liang P, Zhu J, Peng Z, Chao X, Yang Z. Enhanced dielectric performance
Ln-doped BaTiO3 (Ln = La, Ce, Pr, Nd, ~1 × 105 <0.05 [45] of (Ag1/4Nb3/4)0.01Ti0.99O2 ceramic prepared by a wet-chemistry method. Ceram
Sm, Gd, Dy, Ho, and Er) Int 2020;46:11921–5.
[7] Li J, Zeng Y, Fang Y, Chen N, Du G, Zhang A. Synthesis of (La + Nb) co-doped TiO2
rutile nanoparticles and dielectric properties of their derived ceramics composed of
Conclusions submicron-sized grains. Ceram Int 2021;47:8859–67.
[8] Zhu J, Wu D, Liang P, Zhou X, Peng Z, Chao X, et al. Ag+/W6+ co-doped TiO2
ceramic with colossal permittivity and low loss. J Alloy Compd 2021;856:157350.
CoTTO ceramics were synthesized using a mixed oxide method, [9] Liang P, Zhu J, Wu D, Peng H, Chao X, Yang Z. Good dielectric performance and
making the homogeneous dispersion of the Co2+ and Ta5+ dopants. broadband dielectric polarization in Ag, Nb co-doped TiO2. J Am Ceram Soc 2021;
Single phase and dense ceramic samples were obtained by sintering 104:2702–10.
[10] Liu J, Duan C-G, Yin W-G, Mei W, Smith R, Hardy J. Large dielectric constant and
process at 1400 ◦ C for 5 h. The lattice parameters were increased due to Maxwell-Wagner relaxation in Bi2/3Cu3Ti4O12. Phys Rev B 2004;70:144106.
relatively large ionic radii of the co–dopants. Low tanδ of 0.015–0.043 [11] Krohns S, Lunkenheimer P, Ebbinghaus SG, Loidl A. Colossal dielectric constants in
with ultra-high ε′ of 7.0–9.1 × 104 at 1 kHz can be obtained in the single-crystalline and ceramic CaCu3Ti4O12 investigated by broadband dielectric
spectroscopy. J Appl Phys 2008;103:084107.
sintered CoTTO. Moreover, Δε′ ≤ ±15 % can be achieved in the tem­ [12] Liu L, Fan H, Fang P, Chen X. Sol–gel derived CaCu3Ti4O12 ceramics: Synthesis,
perature range of − 60–200 ◦ C. Both defect dipoles and IBLC effect can characterization and electrical properties. Mater Res Bull 2008;43:1800–7.
contribute to the overall dielectric response. The contributions of [13] Rhouma S, Megriche A, El Amrani M, Said S, Roger S, Autret-Lambert C. Effect of
Sr/Mg co-doping on the structural, dielectric, and electrical properties of
extrinsic and intrinsic factors were separated by the dielectric re­ CaCu3Ti4O12 ceramics. J Mater Sci Mater Electron 2022;33(7):4535–49.
laxations in low and high–frequency ranges, respectively. Low tanδ and [14] Liu L, Fan H, Fang P, Jin L. Electrical heterogeneity in CaCu3Ti4O12 ceramics
excellent temperature stability of the DPs were attributed to the IBLC fabricated by sol–gel method. Solid State Commun 2007;142:573–6.
[15] Liu L, Fan H, Chen X, Fang P. Electrical properties and microstructural
effect associated with a high resistance of the insulating part and defect
characteristics of nonstoichiometric CaCu3xTi4O12 ceramics. J Alloy Compd 2009;
dipoles. 469:529–34.
[16] Du G, Wei F, Li W, Chen N. Co-doping effects of A-site Y3+ and B-site Al3+ on the
microstructures and dielectric properties of CaCu3Ti4O12 ceramics. J Eur Ceram
CRediT authorship contribution statement
Soc 2017;37:4653–9.
[17] Peng Z, Zhou X, Wang J, Zhu J, Liang P, Chao X, et al. Origin of colossal
Theeranuch Nachaithong: Conceptualization, Methodology, permittivity and low dielectric loss in Na1/3Cd1/3Y1/3Cu3Ti4O12 ceramics. Ceram
Investigation, Validation, Visualization, Writing – original draft. Nar­ Int 2020;46:11154–9.
[18] Peng Z, Wang J, Zhou X, Zhu J, Lei X, Liang P, et al. Grain engineering inducing
ong Chanlek: Formal analysis, Investigation, Validation. Pairot high energy storage in CdCu3Ti4O12 ceramics. Ceram Int 2020;46:14425–30.
Moontragoon: Formal analysis, Validation. Pornjuk Srepusharawoot: [19] Miao G-T, Wang Z-J, Li P, Hao J-G, Li W, Du J, et al. Complex impedance and
Software, Data curation, Formal analysis, Validation. Prasit Thongbai: electrical conduction analysis of Ho2O3 doped CaCu3Ti4O12 NTC ceramics. J Asian
Ceram Soc 2022;10:165–77.
Conceptualization, Data curation, Methodology, Resources, Validation, [20] Zhao J, Sun L, Cao E, Hao W, Zhang Y, Ju L. Strontium doped CaCu3Ti4O12
Writing – original draft, Project administration, Writing – review & ceramics with very low dielectric loss synthesized by the sol–gel method. Appl Phys
editing, Funding acquisition. A 2022;128:340.
[21] Meng Y, Liu K, Zhang X, Qiang X, Lei X, Chen J, et al. Compositional modulation
and annealing treatment in BaTiO3 to simultaneously achieve colossal permittivity,
Declaration of Competing Interest low dielectric loss, and high thermal stability. Ceram Int 2021;47(23):33912–6.
[22] Wu J, Nan C-W, Lin Y, Deng Y. Giant dielectric permittivity observed in Li and Ti
doped NiO. Phys Rev Lett 2002;89:217601.
The authors declare that they have no known competing financial [23] Li Y, Fang L, Liu L, Huang Y, Hu C. Giant dielectric response and charge
interests or personal relationships that could have appeared to influence compensation of Li- and Co-doped NiO ceramics. Mater Sci Eng B 2012;177(9):
the work reported in this paper. 673–7.
[24] Song Y, Wang X, Zhang X, Sui Y, Zhang Y, Liu Z, et al. The contribution of doped-Al
to the colossal permittivity properties of AlxNb0.03Ti0.97-xO2 rutile ceramics.
Data availability J Mater Chem C 2016;4(28):6798–805.
[25] Tuichai W, Danwittayakul S, Chanlek N, Thongbai P, Maensiri S. High-
performance giant-dielectric properties of rutile TiO2 co-doped with acceptor-Sc3+
Data will be made available on request. and donor-Nb5+ ions. J Alloy Compd 2017;703:139–47.
[26] Tuichai W, Thongyong N, Danwittayakul S, Chanlek N, Srepusharawoot P,
Acknowledgments Thongbai P, et al. Very low dielectric loss and giant dielectric response with
excellent temperature stability of Ga3+ and Ta5+ co-doped rutile-TiO2 ceramics.
Mater Des 2017;123:15–23.
This project is funded by the National Research Council of Thailand [27] Li Z, Wu J, Wu W. Composition dependence of colossal permittivity in
(NRCT): (N41A640084). The research received funding support from (Sm0.5Ta0.5)xTi1-xO2 ceramics. J Mater Chem C 2015;3(35):9206–16.
[28] Li Z, Wu J, Xiao D, Zhu J, Wu W. Colossal permittivity in titanium dioxide ceramics
the Research and Graduate Studies Office of Khon Kaen University. This modified by tantalum and trivalent elements. Acta Mater 2016;103:243–51.
work has received scholarship under the Post-Doctoral Training Pro­ [29] Youssef AM, Yakout SM, Mousa SM. High relative permittivity and excellent dye
gram from Khon Kaen University, Thailand (PD 2565-12). photo-elimination: Pure and (Zr4+, Y3+, Sb5+) multi-doped anatase TiO2 structure.
Opt Mater 2023;135:113261.
[30] Wang L, Liu X, Zhang M, Bi X, Ma Z, Li J, et al. Colossal dielectric behavior of (Nb,
Appendix A. Supplementary data Ga) co-doped TiO2 single crystal. J Alloy Compd 2022;921:166053.

Supplementary data to this article can be found online at https://doi.

6
T. Nachaithong et al. Results in Physics 47 (2023) 106334

[31] Fan J, Long Z, Zhou H, He G, Hu Z. Colossal permittivity of (Tm+Nb) co-doped [43] Parker JC, Siegel RW. Calibration of the Raman spectrum to the oxygen
rutile-TiO2 ceramics with ultralow dielectric loss and excellent thermal stability. stoichiometry of nanophase TiO2. Appl Phys Lett 1990;57(9):943–5.
J Alloy Compd 2022;921:166200. [44] Wei X, Jie W, Yang Z, Zheng F, Zeng H, Liu Y, et al. Colossal permittivity properties
[32] Tuichai W, Danwittayakul S, Maensiri S, Thongbai P. Investigation on temperature of Zn, Nb co-doped TiO2 with different phase structures. J Mater Chem C 2015;3
stability performance of giant permittivity (In + Nb) in co-doped TiO2 ceramic: a (42):11005–10.
crucial aspect for practical electronic applications. RSC Adv 2016;6(7):5582–9. [45] Meng Y, Liu K, Zhang X, Lei X, Chen J, Yang Z, et al. Defect engineering in rare-
[33] Wu YQ, Zhao X, Zhang JL, Su WB, Liu J. Huge low-frequency dielectric response of earth-doped BaTiO3 ceramics: Route to high-temperature stability of colossal
(Nb, In)-doped TiO2 ceramics. Appl Phys Lett 2015;107(24):242904. permittivity. J Am Ceram Soc 2022;105(9):5725–37.
[34] Dong W, Hu W, Berlie A, Lau K, Chen H, Withers RL, et al. Colossal dielectric [46] Liu J, Duan C-G, Mei WN, Smith RW, Hardy JR. Dielectric properties and Maxwell-
behavior of Ga+Nb co-doped rutile TiO2. ACS Appl Mater Interfaces 2015;7(45): Wagner relaxation of compounds ACu3Ti4O12 (A=Ca, Bi2∕3, Y2∕3, La2∕3). J Appl
25321–5. Phys 2005;98:093703.
[35] Shannon RD. Revised effective ionic radii and systematic studies of interatomie [47] Liu J, Sui Y, Duan C-G, Mei W-N, Smith RW, Hardy JR. CaCu3Ti4O12: low-
distances in halides and chaleogenides. Acta Cryst A 1976;32:751–67. temperature synthesis by pyrolysis of an organic solution. Chem Mater 2006;18:
[36] Liu G, Fan H, Xu J, Liu Z, Zhao Y. Colossal permittivity and impedance analysis of 3878–82.
niobium and aluminum co-doped TiO2 ceramics. RSC Adv 2016;6(54):48708–14. [48] Sarkar S, Jana PK, Chaudhuri BK. Colossal internal barrier layer capacitance effect
[37] Wang J, Lu Z, Deng T, Zhong C, Chen Z. Improved dielectric, nonlinear and in polycrystalline copper (II) oxide. Appl Phys Lett 2008;92:022905.
magnetic properties of cobalt-doped CaCu3Ti4O12 ceramics. J Eur Ceram Soc 2018; [49] Liu L, Ren S, Liu J, Han F, Zhang J, Peng B, et al. Localized polarons and conductive
38(10):3505–11. charge carriers: understanding CaCu3Ti4O12 over a broad temperature range. Phys
[38] Dong W, Chen D, Hu W, Frankcombe TJ, Chen H, Zhou C, et al. Colossal Rev B 2019;99:094110.
permittivity behavior and its origin in rutile (Mg1/3Ta2/3)xTi1-xO2. Sci Rep 2017;7: [50] Li G, Chen Z, Sun X, Liu L, Fang L, Elouadi B. Electrical properties of AC3B4O12-type
9950. perovskite ceramics with different cation vacancies. Mater Res Bull 2015;65:
[39] Adams T, Sinclair D, West A. Characterization of grain boundary impedances in 260–5.
fine- and coarse-grained CaCu3Ti4O12 ceramics. Phys Rev B 2006;73:094124. [51] Huang Y, Liu L, Shi D, Wu S, Zheng S, Fang L, et al. Giant dielectric permittivity
[40] Tuichai W, Danwittayakul S, Chanlek N, Thongbai P. Nonlinear current-voltage and non-linear electrical behavior in CaCu3Ti4O12 varistors from the molten-salt
and giant dielectric properties of Al3+ and Ta5+ co-doped TiO2 ceramics. Mater Res synthesized powder. Ceram Int 2013;39:6063–8.
Bull 2019;116:137–42. [52] Hao W, Zhang J, Tan Y, Zhao M, Wang C. Giant dielectric permittivity properties
[41] Brežný B, Muan A. Phase relations and stabilities of compounds in the system CoO and relevant mechanism of NaCu3Ti3SbO12 ceramics. J Am Ceram Soc 2011;94:
TiO2. J Inorg Nucl Chem 1969;31(3):649–55. 1067–72.
[42] Hu W, Lau K, Liu Y, Withers RL, Chen H, Fu L, et al. Colossal dielectric permittivity
in (Nb+Al) codoped rutile TiO2 ceramics: compositional gradient and local
structure. Chem Mater 2015;27(14):4934–42.

You might also like